首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 0 毫秒
1.
Mutation of the mouse Usp14 gene, encoding the homolog of yeast deubiquitinating enzyme Ubp6, causes ataxia. Here we show that deletion of the UBP6 gene in Saccharomyces cerevisiae causes sensitivity to a broad range of toxic compounds and antagonizes phenotypic expression and de novo induction of the yeast prion [PSI+], a functionally defective self-perpetuating isoform of the translation termination factor Sup35. Conversely, overexpression of ubiquitin (Ub) increases phenotypic expression and induction of [PSI+] in the wild type cells and suppresses all tested ubp6Delta defects, indicating that they are primarily due to depletion of cellular Ub levels. Several lines of evidence suggest that Ubp6 functions on the proteasome. First, Ub levels in the ubp6Delta cells can be partly restored by proteasome inhibitors, suggesting that deletion of Ubp6 decreases Ub levels by increasing proteasome-dependent degradation of Ub. Second, fluorescence microscopy analysis shows that Ubp6-GFP fusion protein is localized to the nucleus of yeast cell, as are most proteasomes. Third, the N-terminal Ub-like domain, although it is not required for nuclear localization of Ubp6, targets Ubp6 to the proteasome and cannot be functionally replaced by Ub. The human ortholog of Ubp6, USP14, probably plays a similar role in higher eukaryotes, since it fully compensates for ubp6Delta defects and binds to the yeast proteasome. These data link the Ub system to prion expression and propagation and have broad implications for other neuronal inclusion body diseases.  相似文献   

2.
《朊病毒》2013,7(2):60-65
Yeast prions provide a powerful model system for examining prion formation and propagation in vivo. Yeast prion formation is driven primarily by amino acid composition, not by primary amino acid sequence. However, although yeast prion domains are consistently glutamine/asparagine-rich, they otherwise vary significantly in their compositions. Therefore, elucidating the exact compositional requirements for yeast prion formation has proven challenging. We have developed an in vivo method that allows for estimation of the prion propensity of each amino acid within the context of a yeast prion domain.1 Using these values, we are able to predict the prion-propensity of various glutamine/asparagine-rich yeast domains. These results provide insight into the basis for yeast prion formation, and may aid in the discovery of additional novel prion domains. Additionally, we examined whether amino acid composition could drive interactions between heterologous glutamine/asparagine-rich proteins.2 Although inefficient interactions between yeast prion domains have previously been observed, we found that one prion protein, Ure2, is able to interact with compositionally similar domains with unprecedented efficiency. This observation, combined with the growing number of yeast prions, suggests that a broad network of interactions between heterologous glutamine/asparagine-rich proteins may affect yeast prion formation.  相似文献   

3.
Yeast prions provide a powerful model system for examining prion formation and propagation in vivo. Yeast prion formation is driven primarily by amino acid composition, not by primary amino acid sequence. However, although yeast prion domains are consistently glutamine/asparagine-rich, they otherwise vary significantly in their compositions. Therefore, elucidating the exact compositional requirements for yeast prion formation has proven challenging. We have developed an in vivo method that allows for estimation of the prion propensity of each amino acid within the context of a yeast prion domain.1 Using these values, we are able to predict the prion-propensity of various glutamine/asparagine-rich yeast domains. These results provide insight into the basis for yeast prion formation, and may aid in the discovery of additional novel prion domains. Additionally, we examined whether amino acid composition could drive interactions between heterologous glutamine/asparagine-rich proteins.2 Although inefficient interactions between yeast prion domains have previously been observed, we found that one prion protein, Ure2, is able to interact with compositionally similar domains with unprecedented efficiency. This observation, combined with the growing number of yeast prions, suggests that a broad network of interactions between heterologous glutamine/asparagine-rich proteins may affect yeast prion formation.Key words: yeast, prion, amyloid, Ure2, Sup35The highly-studied yeast prion proteins Sup35, Ure2 and Rnq1, which form the [PSI+], [URE3] and [PIN+] prions, respectively, each contain a glutamine/asparagine (Q/N) rich domain that drives prion formation. 3 Prion formation is thought to involve conversion of the soluble proteins into an insoluble amyloid form.3 Four additional yeast prion proteins containing Q/N-rich prion-forming domains have recently been discovered, and domains with similar Q/N-content are over-represented in various eukaryotic genomes.48 This suggests that numerous other prion proteins may remain to be discovered, yet predicting which of these Q/N-rich domains is likely to form prions has proven difficult.Randomizing the order of the amino acids in either the Sup35 or Ure2 prion domains does not block prion formation, demonstrating that prion formation is driven primarily by amino acid composition, not primary sequence.9,10 We therefore explored two related questions. First, can amino acid composition be used to accurately predict prion propensity?1 Second, to what extent can compositionally similar Q/N-rich proteins interact during prion formation?2Surprisingly, although composition drives yeast prion formation, compositional similarity to known prion domains is a poor predictor of prion propensity. In addition to high Q/N content, yeast prion domains show an under-representation of both charged and hydrophobic residues (1,11 Alberti et al. recently identified the 100 yeast domains with greatest compositional similarity to the known prion domains, and tested each domain in four different assays for prion-like activity.4 This remarkable effort revealed numerous new potential prion domains; however, there was very little correlation between a domain’s compositional similarity to known prion domains and its prion-like activity.1,4 The most likely explanation for this unexpected finding is that because the yeast prion domains are likely not optimized for maximum prion propensity, some compositional deviations will increase prion propensity and some will decrease it. However, algorithms that predict prion propensity based on compositional similarity to known prion domains are predicated on the assumption that all compositional changes will reduce prion propensity. Instead, accurate prediction of prion propensity requires understanding how changes in amino acid composition affect prion propensity.

Table 1

Prion propensity, order propensity and prevalence for each amino acid
Amino acidPrion propensityaPrevalence in prion domainsbHydrophobicityc
Phenylalanine0.842.50.81
Isoleucine0.811.11.0
Valine0.811.50.97
Tyrosine0.787.80.36
Methionine0.671.70.71
Tryptophan0.6700.40
Cysteine0.4200.78
Serine0.1310.00.41
Asparagine0.08023.50.11
Glutamine0.06922.00.11
Glycine−0.03912.90.46
Leucine−0.0402.30.92
Threonine−0.122.10.42
Histidine−0.280.90.14
Alanine−0.403.90.70
Arganine−0.412.50.00
Glutamic Acid−0.611.50.11
Proline−1.201.70.32
Aspartic Acid−1.281.30.11
Lysine−1.580.70.07
Open in a separate windowaExperimentally determined prion propensity from ref. 1. Prion propensities were calculated as the natural log of the fold over/under-representation of the amino acid among prion-forming clones relative to the naïve library.bThe average percent prevalence among the Sup35, rnq1 and Ure2 prion domains.cFrom ref. 21, rescaled from 0 to 1. This scale is used by FoldIndex to calculate intrinsic folding propensity (IF) by the equation IF = 2.785 — <R> — 1.151, where <H> is the mean hydrophobicity and <R> is average net charge.20We therefore developed the first method to quantify the prion propensity of each amino acid within the context of a Q/N-rich prion domain (Fig. 1). We utilized Sup35-27, a scrambled version of Sup35 that forms prions with high efficiency. Using an oligonucleotide-based mutagenesis approach, we replaced eight consecutive codons within the SUP35-27 DNA sequence with (NNB)8, where N represents any of the four nucleotides and B represents any nucleotide except adenine. This generated a library of sequences in which all 20 amino acids should be present at each position within the mutated region. We then selected for the subset of proteins that maintained the ability to form prions. By comparing the naïve library to the prion-forming subset, we were able to determine the degree to which each amino acid was over- or under-represented among the prion-forming clones. These numbers were then used to generate a scale ranking the prion-propensity of all 20 amino acids (1,11,12 By contrast, hydrophobic residues were strongly overrepresented, and no detectable bias was seen for or against glutamines and asparagines.1 Similar results were seen at a second position, demonstrating that these biases were not an artifact of local interactions.Open in a separate windowFigure 1Mutagenesis method. Oligonucleotides were designed with a degenerate region, flanked by regions complementary to SUP35-27. These degenerate oligonucleotides were used to PCR amplify SUP35-27, generating a library of sequences in which eight codons were randomly mutated. Using plasmid shuffling, the library was used to replace wild-type Sup35 as the sole copy of Sup35 in the cell. Mutants were selected for prion formation by plating on medium lacking adenine. Isolates from the naïve library and the prion-forming subset were sequenced to determine which amino acids were over/under-represented among the prion-forming isolates.Many of these results can be rationalized based on the proposed structure of Sup35-27 amyloid fibrils. Amyloid fibrils are β-sheet-rich structures in which the β-strands are predominantly oriented perpendicular to the long axis of the fibril. Solid state NMR indicates that for Sup35-27, the β-strands adopt an inregister parallel alignment.13 This inregister alignment should strongly disfavor charged residues due to electrostatic repulsion. Proline residues are known β-sheet breakers, so it is not surprising that they would also be disfavored. Hydrophobic residues have been proposed to drive formation of other inregister parallel amyloid structures, so their high prion-propensities in our assay should not be a complete surprise.14Other aspects of our results are less easily explained based on proposed structures. X-ray diffraction studies of small fragments from Sup35 indicated that hydrogen bonding of in-register glutamines and asparagines stabilizes prion fibers, analogous to the polar zippers that have been proposed for poly-Q.15,16 However, we saw no bias in favor of glutamines and asparagines. This lack of bias in favor of glutamines and asparagines was particularly surprising based on the compositions of known yeast prion domains. Although high Q/N content is not an absolute requirement for a protein to act as a prion in yeast, Q/N residues are consistently over-represented in yeast prion domains (11 Likewise, although we observed a strong bias in favor of hydrophobic residues, and although tyrosines have been implicated in prion formation and propagation, hydrophobic residues in general are strongly under-represented among yeast prion domains (1,11,17,18This raises the question, why is there such a large discrepancy between the residues that most strongly promote prion formation and those that are actually present in yeast prion domains? The simplest explanation is that the composition of yeast prion domains may reflect functions of the domains other than prion formation. Bioinformatic analysis can reveal what compositional features are present in prion domain, but not why they are present. Nevertheless, it seems surprising that yeast prion domains could form prions when they are largely lacking in the most strongly prion-promoting residues. We propose that native state intrinsic disorder is essential to yeast prion formation, and can help explain the ability of yeast prion domains to form prions despite their relative lack of strongly prion-prone amino acids. The yeast prion domains are intrinsically disordered, and for Sup35, this structural flexibility seems to be important for prion formation.19 The yeast prion domains are biased towards residues that balance intrinsic disorder and prion-propensity. For most proteins, amyloid formation must compete with native-state structure; in the absence of this competition, yeast prion domains are able to form prions despite relatively modest prion propensities. Although hydrophobic residues have high prion propensity, large numbers of hydrophobic residues would reduce intrinsic disorder by promoting hydrophobic collapse, potentially creating a stable fold that would compete with amyloid formation.A significant question was whether mutagenesis results from short eightresidue segments could be used to predict prion propensities of entire prion domains. To test this, we scanned various proteins with a 41-amino-acid window size, calculating for each window prion propensity as the sum of the experimentally determined prion propensities for each amino acid in the window, and the order propensity using FoldIndex.20 FoldIndex is a simple algorithm for identifying disordered regions based on hydrophobicity (using the Kyte/Doolittle scale) and net charge.20,21 Strikingly, regions with similar predicted prion propensity to the yeast prion domains are common in non-prion proteins; however, these regions are consistently predicted to be ordered.1 By contrast, for Sup35 and Ure2, the prion domains are easily distinguishable as regions with positive prion propensity and negative FoldIndex order propensity (Fig. 2), supporting a role for intrinsic disorder in facilitating prion formation. Although the prion domains for the other yeast prion proteins are not as clearly defined, similar disordered and prion-prone regions are found in all of the yeast prion domains except that of Cyc8. By contrast, while Q/N-rich domains with low prion-like activity are consistently predicted to be disordered, they generally have low predicted prion propensity. By scoring proteins based on the predicted prion propensity of the most prion-prone region, we were able to distinguish with greater than 90% accuracy between Q/N-rich domains with and without prion-like activity. This is the first time that such prediction accuracy has been achieved for Q/N-rich domains.Open in a separate windowFigure 2Prion propensity maps. Ure2p (A) and Sup35 (B) were scanned using a window size of 41 amino acids, calculating for each window the average order propensity using FoldIndex and prion propensity as the sum of the experimentally determined prion propensities for each amino acid across the window. The prion domain (PFD) is shaded.These results provide insight into the differences between Q/N-rich and non-Q/N-rich amyloid proteins. Amyloid formation by non-Q/N-rich proteins is thought to be driven by short peptide stretches.22 Therefore, most amyloid prediction algorithms are designed to look for local regions of high prion propensity. However, our algorithm uses a relatively large 41-amino-acid window size. When we used a smaller window size, our algorithm lost the ability to distinguish between Q/N-rich domains with and without prion-like activity. Thus, the yeast prion domains are characterized by extended disordered regions of modest prion propensity, not local regions of high prion propensity. This explains why many amyloid prediction algorithms underestimate the amyloid propensity of the polar yeast prion domains.23An obvious question is whether mutagenesis data from a single protein can really be used to make such broad claims about the basis for prion formation by all Q/N-rich yeast prion proteins. Similarly, FoldIndex is a relatively simple disorder-prediction algorithm that may not be specifically optimized for Q/N-rich proteins. However, these concerns are somewhat allayed by the accurate predictions that our algorithm, when combined with FoldIndex, is able to make. The fact that mutagenesis data from a single Q/N-rich protein allows for accurate prediction of a broad range of other Q/N-rich proteins suggests that the basic mechanisms of prion formation are similar for the various Q/N-rich prion domains. There is little doubt that a broader data set derived from multiple yeast prion domains could improve prediction accuracy; likewise, other disorder-prediction algorithms might be better suited for analyzing yeast prions. Nevertheless, the prediction accuracy that we have achieved suggests that while such changes might modestly improve our prediction accuracy, they are unlikely to fundamentally change our general conclusions about yeast prion domains.Although our results provide an exciting first step towards understanding the sequence requirements for prion formation, numerous questions remain unanswered. For Sup35, separate regions of the prion domain are required for prion formation and prion propagation.24 Therefore, future experiments will be needed to dissect the distinct sequence requirements for prion formation versus propagation. Likewise, although numerous domains have been identified that can drive prion-like activity when inserted in the place of the Sup35 prion domain,4,25 flanking sequences outside of the Sup35 prion domain affect prion propagation.26 Therefore, there is no guarantee that such domains will be able to act as prions in their native contexts. Understanding the role of flanking sequences will be critical for predicting prion propensity of domains within their native contexts. Similarly, expression levels and patterns and cellular localization likely influence prion formation and propagation.The effects of primary sequence on prion propensity are also still unclear. The dominant effect of amino acid composition on prion formation has made it difficult to distinguish the more subtle effects of primary sequence. By delineating the compositional requirements for prion formation, our work should make it easier to identify primary sequence elements that affect prion propensity. Specifically, examination of outlier proteins not accurately predicted by our composition-based algorithm may reveal primary sequence features that promote or inhibit prion formation.One such aspect of primary sequence is already incorporated into our algorithm. In examining the 100 Q/N-rich proteins studied by Alberti et al. we discovered a subtle, but statistically significant bias in the distribution of proline residues.1,4 When prolines were present in Q/N-rich domains with prion-like activity, they tended to occur in clusters, while the prolines were more likely to be dispersed in domains without prion-like activity. This is not surprising, as prolines are known β-sheet breakers. If multiple prolines are present in single cluster, they will disrupt the β-sheet structure at a single location; by contrast, the same number of prolines dispersed throughout a sequence will result in multiple locations where the β-sheet structure is disrupted.Another potential example of the effects of primary sequence is seen in Cyc8, the only yeast prion protein incorrectly predicted by our algorithm not to form prions. Interestingly, the prion domain contains an imperfect (QA)32 repeat. Patterns of alternating polar and non-polar residues are thought to promote amyloid formation, although alanine has not generally been included among the non-polar residues when considering such patterns.27 Therefore, this primary sequence element may explain why our algorithm, which predominantly considers amino acid composition, predicts Cyc8 to have relatively low prion propensity.The high level of prediction accuracy achieved by our algorithm should also facilitate the identification of new prion proteins. The large number of yeast prion proteins, combined with the prevalence of Q/N-rich domains in eukaryotic genomes, suggests that similar prion-like structural conversions may be common in other organisms. However, only two prion proteins have been identified outside of Saccharomyces cerevisiae—the mammalian protein PrP and Podospora anserine protein HET-s—neither of which is Q/N-rich. Identification of new prion proteins has been hindered by the lack of accurate prediction algorithms. In yeast, genetic screens and compositional homology searches have led to the identification of novel prions, but with relatively low success rates.48 Understanding the compositional requirements for prion formation should improve this success rate, allowing for more targeted testing of potential prion proteins.As the list of Q/N-rich yeast prion proteins continues to grow, an obvious question will be, how do interactions between heterologous prion proteins affect prion formation and propagation? If amino acid composition is the predominant factor determining prion-like activity, can amino acid composition also drive interactions between heterologous prion proteins? Various interactions have been observed among the yeast prion proteins. Under normal cellular conditions, prion formation by Sup35 requires the presence of [PIN+].28 In vitro and in vivo evidence suggests that Rnq1 aggregates can act as imperfect templates for seeding Sup35 aggregation.29 Overexpression of various Q/N-rich proteins can substitute for [PIN+], allowing [PSI+] formation in cells lacking [PIN+]; this suggests that Sup35 can interact with a broad range of Q/N-rich proteins.30,31 [PIN+] also promotes [URE3] formation, while [PSI+] inhibits [URE3] formation.32,33 However, each of these interactions is quite inefficient; for example, Rnq1 aggregates are at least 50-fold less efficient than Sup35 aggregates at seeding Sup35 amyloid formation.29Therefore, we were surprised to discover that [URE3] prion formation can be stimulated with high efficiency by a variety of compositionally similar domains.2 Overexpression of the prion domain of Ure2 significantly increases [URE3] formation. Surprisingly, overexpression of each of the scrambled versions of Ure2 stimulates wild-type [URE3] formation with comparable efficiency.2 In vitro, amyloid aggregates formed by the scrambled Ure2s were able to efficiently seed amyloid aggregation by wild-type Ure2, suggesting that wild-type Ure2 is able to directly interact with scrambled Ure2 amyloid aggregates. There were limits to Ure2’s promiscuity, as overexpression of scrambled Sup35 prion domains did not promote [URE3] formation.To determine whether Ure2 could similarly interact with naturally occurring yeast proteins, we performed a genomic search to identify the five yeast protein fragments with greatest compositional similarity to the Ure2 prion domain.2 Four of these five fragments were able to stimulate wild-type [URE3] formation. The efficiency of this cross-seeding was unprecedented; one of the fragments, from Sap30, was able to stimulate [URE3] formation three-fold more efficiently than the wild-type Ure2 prion domain. These results raise the possibility that such interactions could be physiologically relevant, acting as regulators of either beneficial or deleterious amyloid formation. Interestingly, not all prion proteins appear to be as promiscuous as Ure2. Prion formation by Sup35 was not significantly affected by overexpression of scrambled versions of Sup35. Defining the sequence features that allow Ure2 to interact with compositionally similar domains, and defining the limits of Ure2’s promiscuity, will be critical for determining the extent to which similar interactions affect the normal physiology of yeast prions and mammalian amyloid diseases.Although it is clear that Ure2 is able to promiscuously interact with a broad range of compositionally similar proteins, our results say nothing about why such interactions occur. These interactions may have evolved to positively or negatively regulate prion formation. Alternatively, such prion-promoting interactions may simply be a byproduct of the natural function of the Ure2 prion domain. Although the exact function of the Ure2 prion domain is unknown, regions of intrinsic disorder domains are often used to recognize multiple binding partners.34 Promiscuous prion-promoting interactions may simply be an unintended consequence of such activity. Therefore, future studies will be needed to determine the function, if any, of Ure2’s promiscuity. Likewise, although our mutagenesis results clearly indicate that the compositions of the yeast prion domains reflect a balance of intrinsic disorder and prion propensity, our studies say nothing about why the yeast prion domains have evolved this balance. These compositional characteristics may have evolved for a reason unrelated to prion formation. Alternatively, the yeast prion domains may have specifically been optimized for the purpose of prion formation.  相似文献   

4.
A variety of proteins are capable of converting from their soluble forms into highly ordered fibrous cross‐β aggregates (amyloids). This conversion is associated with certain pathological conditions in mammals, such as Alzheimer disease, and provides a basis for the infectious or hereditary protein isoforms (prions), causing neurodegenerative disorders in mammals and controlling heritable phenotypes in yeast. The N‐proximal region of the yeast prion protein Sup35 (Sup35NM) is frequently used as a model system for amyloid conversion studies in vitro. Traditionally, amyloids are recognized by their ability to bind Congo Red dye specific to β‐sheet rich structures. However, methods for quantifying amyloid fibril formation thus far were based on measurements linking Congo Red absorbance to concentration of insulin fibrils and may not be directly applicable to other amyloid‐forming proteins. Here, we present a corrected formula for measuring amyloid formation of Sup35NM by Congo Red assay. By utilizing this corrected procedure, we explore the effect of different sodium salts on the lag time and maximum rate of amyloid formation by Sup35NM. We find that increased kosmotropicity promotes amyloid polymerization in accordance with the Hofmeister series. In contrast, chaotropes inhibit polymerization, with the strength of inhibition correlating with the B‐viscosity coefficient of the Jones‐Dole equation, an increasingly accepted measure for the quantification of the Hofmeister series.  相似文献   

5.
The number of known fungal proteins capable of switching between alternative stable conformations is steadily increasing, suggesting that a prion-like mechanism may be broadly utilized as a means to propagate altered cellular states. To gain insight into the mechanisms by which cells regulate prion formation and toxicity we examined the role of the yeast ribosome-associated complex (RAC) in modulating both the formation of the [PSI+] prion – an alternative conformer of Sup35 protein – and the toxicity of aggregation-prone polypeptides. The Hsp40 RAC chaperone Zuo1 anchors the RAC to ribosomes and stimulates the ATPase activity of the Hsp70 chaperone Ssb. We found that cells lacking Zuo1 are sensitive to over-expression of some aggregation-prone proteins, including the Sup35 prion domain, suggesting that co-translational protein misfolding increases in Δzuo1 strains. Consistent with this finding, Δzuo1 cells exhibit higher frequencies of spontaneous and induced prion formation. Cells expressing mutant forms of Zuo1 lacking either a C-terminal charged region required for ribosome association, or the J-domain responsible for Ssb ATPase stimulation, exhibit similarly high frequencies of prion formation. Our findings are consistent with a role for the RAC in chaperoning nascent Sup35 to regulate folding of the N-terminal prion domain as it emerges from the ribosome.  相似文献   

6.
《朊病毒》2013,7(2):144-164
Abstract

The number of known fungal proteins capable of switching between alternative stable conformations is steadily increasing, suggesting that a prion-like mechanism may be broadly utilized as a means to propagate altered cellular states. To gain insight into the mechanisms by which cells regulate prion formation and toxicity we examined the role of the yeast ribosome-associated complex (RAC) in modulating both the formation of the [PSI+] prion – an alternative conformer of Sup35 protein – and the toxicity of aggregation-prone polypeptides. The Hsp40 RAC chaperone Zuo1 anchors the RAC to ribosomes and stimulates the ATPase activity of the Hsp70 chaperone Ssb. We found that cells lacking Zuo1 are sensitive to over-expression of some aggregation-prone proteins, including the Sup35 prion domain, suggesting that co-translational protein misfolding increases in Δzuo1 strains. Consistent with this finding, Δzuo1 cells exhibit higher frequencies of spontaneous and induced prion formation. Cells expressing mutant forms of Zuo1 lacking either a C-terminal charged region required for ribosome association, or the J-domain responsible for Ssb ATPase stimulation, exhibit similarly high frequencies of prion formation. Our findings are consistent with a role for the RAC in chaperoning nascent Sup35 to regulate folding of the N-terminal prion domain as it emerges from the ribosome.  相似文献   

7.
8.
The [PSI(+)] prion in yeast has been shown to improve short-term growth in some environments, but its effects on rates of adaptation have not been assessed before now. We adapted three yeast genotypes to three novel environments in the presence and the absence of the prion. There were significant differences in adaptation rates between lines with different combinations of genotype, environment, and prion status. We saw no consistent effect, however, of the prion on the rate of adaptation to new environments. A major factor affecting the rate of adaptation was initial fitness in the new environment: lines with low initial fitness evolved faster than lines with high initial fitness.  相似文献   

9.
Mammalian and fungal prion proteins form self-perpetuating β-sheet-rich fibrillar aggregates called amyloid. Prion inheritance is based on propagation of the regularly oriented amyloid structures of the prion proteins. All yeast prion proteins identified thus far contain aggregation-prone glutamine/asparagine (Gln/Asn)-rich domains, although the mammalian prion protein and fungal prion protein HET-s do not contain such sequences. In order to fill this gap, we searched for novel yeast prion proteins lacking Gln/Asn-rich domains via a genome-wide screen based on cross-seeding between two heterologous proteins and identified Mod5, a yeast tRNA isopentenyltransferase, as a novel non-Gln/Asn-rich yeast prion protein. Mod5 formed self-propagating amyloid fibers in vitro and the introduction of Mod5 amyloids into non-prion yeast induced dominantly and cytoplasmically heritable prion state [MOD+], which harbors aggregates of endogenous Mod5. [MOD+] yeast showed an increased level of membrane lipid ergosterol and acquired resistance to antifungal agents. Importantly, enhanced de novo formation of [MOD+] was observed when non-prion yeast was grown under selective pressures from antifungal drugs. Our findings expand the family of yeast prions to non-Gln/Asn-rich proteins and reveal the acquisition of a fitness advantage for cell survival through active prion conversion.  相似文献   

10.
11.
12.
Self-perpetuating protein aggregates transmit prion diseases in mammals and heritable traits in yeast. De novo prion formation can be induced by transient overproduction of the corresponding prion-forming protein or its prion domain. Here, we demonstrate that the yeast prion protein Sup35 interacts with various proteins of the actin cortical cytoskeleton that are involved in endocytosis. Sup35-derived aggregates, generated in the process of prion induction, are associated with the components of the endocytic/vacuolar pathway. Mutational alterations of the cortical actin cytoskeleton decrease aggregation of overproduced Sup35 and de novo prion induction and increase prion-related toxicity in yeast. Deletion of the gene coding for the actin assembly protein Sla2 is lethal in cells containing the prion isoforms of both Sup35 and Rnq1 proteins simultaneously. Our data are consistent with a model in which cytoskeletal structures provide a scaffold for generation of large aggregates, resembling mammalian aggresomes. These aggregates promote prion formation. Moreover, it appears that the actin cytoskeleton also plays a certain role in counteracting the toxicity of the overproduced potentially aggregating proteins.  相似文献   

13.
Ure2, a regulator of nitrogen metabolism, is the protein determinant of the [URE3] prion state in Saccharomyces cerevisiae. Upon conversion into the prion form, Ure2 undergoes a heritable conformational change to an amyloid-like aggregated state and loses its regulatory function. A number of molecular chaperones have been found to affect the prion properties of Ure2. The studies carried out in our laboratory have been aimed at elucidating the structure of Ure2 fibrils, the mechanism of amyloid formation and the effect of chaperones on the fibril formation of Ure2.  相似文献   

14.
Prions are self-propagating, infectious proteins that underlie several neurodegenerative diseases. The molecular basis underlying their sporadic formation is poorly understood. We show that autophagy protects against de novo formation of [PSI+], which is the prion form of the yeast Sup35 translation termination factor. Autophagy is a cellular degradation system, and preventing autophagy by mutating its core components elevates the frequency of spontaneous [PSI+] formation. Conversely, increasing autophagic flux by treating cells with the polyamine spermidine suppresses prion formation in mutants that normally show a high frequency of de novo prion formation. Autophagy also protects against the de novo formation of another prion, namely the Rnq1/[PIN+] prion, which is not related in sequence to the Sup35/[PSI+] prion. We show that growth under anaerobic conditions in the absence of molecular oxygen abrogates Sup35 protein damage and suppresses the high frequency of [PSI+] formation in an autophagy mutant. Autophagy therefore normally functions to remove oxidatively damaged Sup35, which accumulates in cells grown under aerobic conditions, but in the absence of autophagy, damaged/misfolded Sup35 undergoes structural transitions favoring its conversion to the propagatable [PSI+] form.  相似文献   

15.
16.
Molecular basis of a yeast prion species barrier   总被引:22,自引:0,他引:22  
Santoso A  Chien P  Osherovich LZ  Weissman JS 《Cell》2000,100(2):277-288
The yeast [PSI+] factor is inherited by a prion mechanism involving self-propagating Sup35p aggregates. We find that Sup35p prion function is conserved among distantly related yeasts. As with mammalian prions, a species barrier inhibits prion induction between Sup35p from different yeast species. This barrier is faithfully reproduced in vitro where, remarkably, ongoing polymerization of one Sup35p species does not affect conversion of another. Chimeric analysis identifies a short domain sufficient to allow foreign Sup35p to cross this barrier. These observations argue that the species barrier results from specificity in the growing aggregate, mediated by a well-defined epitope on the amyloid surface and, together with our identification of a novel yeast prion domain, show that multiple prion-based heritable states can propagate independently within one cell.  相似文献   

17.
The dimeric yeast protein Ure2 shows prion-like behaviour in vivo and forms amyloid fibrils in vitro. A dimeric intermediate is populated transiently during refolding and is apparently stabilized at lower pH, conditions suggested to favour Ure2 fibril formation. Here we present a quantitative analysis of the effect of pH on the thermodynamic stability of Ure2 in Tris and phosphate buffers over a 100-fold protein concentration range. We find that equilibrium denaturation is best described by a three-state model via a dimeric intermediate, even under conditions where the transition appears two-state by multiple structural probes. The free energy for complete unfolding and dissociation of Ure2 is up to 50 kcal mol(-1). Of this, at least 20 kcal mol(-1) is contributed by inter-subunit interactions. Hence the native dimer and dimeric intermediate are significantly more stable than either of their monomeric counterparts. The previously observed kinetic unfolding intermediate is suggested to represent the dissociated native-like monomer. The native state is stabilized with respect to the dimeric intermediate at higher pH and in Tris buffer, without significantly affecting the dissociation equilibrium. The effects of pH, buffer, protein concentration and temperature on the kinetics of amyloid formation were quantified by monitoring thioflavin T fluorescence. The lag time decreases with increasing protein concentration and fibril formation shows pseudo-first order kinetics, consistent with a nucleated assembly mechanism. In Tris buffer the lag time is increased, suggesting that stabilization of the native state disfavours amyloid nucleation.  相似文献   

18.
朊病毒对中枢神经系统的影响及其作用机制   总被引:4,自引:1,他引:3  
本文着重叙述了以几个问题:(1)朊病毒能否从牛转移感染到人;(2)朊病毒的感感染途径是怎样的;(3)朊病毒感染有何条件;(4)正常的朊病毒蛋白在神经元和胶质细胞表面表达有何重要功能;(5)朊病毒蛋白功能的结构基因;(6)朊病毒的可能作用机制。最后,分析了朊病毒影响中枢神经系统的两条可能途径,即破坏正常朊病毒蛋白的功能和与小神经胶质细胞共同作用提高神经元对自由基的敏感性及增加氧化自由基的含量。  相似文献   

19.
The frequency with which the yeast [PSI(+)] prion form of Sup35 arises de novo is controlled by a number of genetic and environmental factors. We have previously shown that in cells lacking the antioxidant peroxiredoxin proteins Tsa1 and Tsa2, the frequency of de novo formation of [PSI(+)] is greatly elevated. We show here that Tsa1/Tsa2 also function to suppress the formation of the [PIN(+)] prion form of Rnq1. However, although oxidative stress increases the de novo formation of both [PIN(+)] and [PSI(+)], it does not overcome the requirement of cells being [PIN(+)] to form the [PSI(+)] prion. We use an anti-methionine sulfoxide antibody to show that methionine oxidation is elevated in Sup35 during oxidative stress conditions. Abrogating Sup35 methionine oxidation by overexpressing methionine sulfoxide reductase (MSRA) prevents [PSI(+)] formation, indicating that Sup35 oxidation may underlie the switch from a soluble to an aggregated form of Sup35. In contrast, we were unable to detect methionine oxidation of Rnq1, and MSRA overexpression did not affect [PIN(+)] formation in a tsa1 tsa2 mutant. The molecular basis of how yeast and mammalian prions form infectious amyloid-like structures de novo is poorly understood. Our data suggest a causal link between Sup35 protein oxidation and de novo [PSI(+)] prion formation.  相似文献   

20.
Prion diseases propagate by converting a normal glycoprotein of the host, PrP(C), into a pathogenic "prion" conformation. Several misfolding mutants of PrP(C) are degraded through the ER-associated degradation (ERAD)-proteasome pathway. In their infectious form, prion diseases such as bovine spongiform encephalopathy involve PrP(C) of wild-type sequence. In contrast to mutant PrP, wild-type PrP(C) was hitherto thought to be stable in the ER and thus immune to ERAD. Using proteasome inhibitors, we now show that approximately 10% of nascent PrP(C) molecules are diverted into the ERAD pathway. Cells incubated with N-acetyl-leucinal-leucinal-norleucinal (ALLN), lactacystin or MG132 accumulated both detergent-soluble and insoluble PrP species. The insoluble fraction included an unglycosylated 26 kDa PrP species with a protease-resistant core, and a M(r) "ladder" that contained ubiquitylated PrP. Our results show for the first time that wild-type PrP(C) molecules are subjected to ERAD, in the course of which they are dislocated into the cytosol and ubiquitylated. The presence of wild-type PrP molecules in the cytosol may have potential pathogenic implications.  相似文献   

设为首页 | 免责声明 | 关于勤云 | 加入收藏

Copyright©北京勤云科技发展有限公司  京ICP备09084417号